You are on page 1of 21

corrosion and

materials degradation

Review
Corrosion of Carbon Steel in Marine Environments:
Role of the Corrosion Product Layer
Philippe Refait 1, *, Anne-Marie Grolleau 2 , Marc Jeannin 1 , Celine Rémazeilles 1 and René Sabot 1
1 Laboratoire des Sciences de l’Ingénieur pour l’Environnement (LaSIE), UMR 7356 CNRS-La Rochelle
Université, Av. Michel Crépeau, CEDEX 01, F-17042 La Rochelle, France; mjeannin@univ-lr.fr (M.J.);
celine.remazeilles@univ-lr.fr (C.R.); rsabot@univ-lr.fr (R.S.)
2 Naval Group Research, BP 440, CEDEX 50104 Cherbourg-Octeville, France;
anne-marie.grolleau@naval-group.com
* Correspondence: prefait@univ-lr.fr; Tel.: +33-5-46-45-82-27

Received: 5 May 2020; Accepted: 30 May 2020; Published: 3 June 2020 

Abstract: This article presents a synthesis of recent studies focused on the corrosion product layers
forming on carbon steel in natural seawater and the link between the composition of these layers and
the corrosion mechanisms. Additional new experimental results are also presented to enlighten some
important points. First, the composition and stratification of the layers produced by uniform corrosion
are described. A focus is made on the mechanism of formation of the sulfate green rust because this
compound is the first solid phase to precipitate from the dissolved species produced by the corrosion
of the steel surface. Secondly, localized corrosion processes are discussed. In any case, they involve
galvanic couplings between anodic and cathodic zones of the metal surface and are often associated
with heterogeneous corrosion product layers. The variations of the composition of these layers
with the anodic/cathodic character of the underlying metal surface, and in particular the changes in
magnetite content, are thoroughly described and analyzed to enlighten the self-sustaining ability
of the process. Finally, corrosion product layers formed on permanently immersed steel surfaces
were exposed to air. Their drying and oxidation induced the formation of akaganeite, a common
product of marine atmospheric corrosion that was, however, not detected on the steel surface after
the permanent immersion period.

Keywords: carbon steel; seawater; localized corrosion; magnetite; green rust; iron sulfide

1. Introduction
Seawater is one of the most complex and aggressive media. Marine corrosion depends on
numerous interdependent parameters and combines chemical, biological and mechanical factors.
The understanding of the influence of each of these parameters and factors is the key to the optimization
of the design of metal structures and devices used in marine environments. It is also the key to the
optimization of anticorrosion methods and materials performance. Localized corrosion is a particularly
insidious degradation phenomenon and thus a hazard for metal structures integrity. Severe localized
degradations may induce major industrial failures while consuming a very small amount of materials.
In seawater, owing to various combined factors such as materials heterogeneity (grain boundaries,
inclusions, welds . . . ) [1–5], differential aeration [6–9], and biological activity [10–18], the corrosion
processes of carbon steels, though commonly acknowledged as being mainly uniform, are often
localized. The well-known ALWC (Accelerated Low Water Corrosion) phenomenon, which combines
differential aeration and bacteria consortium activity [15,17], perfectly illustrates this point.
Carbon steels are widely used materials for marine applications. They are massively produced
(~1.8 × 109 tons worldwide in 2018 [19]) and are actually inexpensive while they offer good mechanical

Corros. Mater. Degrad. 2020, 1, 198–218; doi:10.3390/cmd1010010 www.mdpi.com/journal/cmd


Corros. Mater. Degrad. 2020, 1 199

properties. They also mainly suffer uniform corrosion and were studied for decades so that their
corrosion rate can be modeled and their lifetime predicted with rather good accuracy in various
marine environments. The phenomenological model proposed by Melchers et al. [20–22] was designed
via a thorough analysis of numerous data. It is based on the description and understanding of the
mechanisms of the uniform corrosion process of iron-based materials. It aims to predict the short and
long-term behavior of steel structures permanently immersed in seawater and forecast reliably the
lifetime of these structures.
Microbiologically influenced corrosion (MIC) is often considered the main reason for the
non-uniformity of carbon steel corrosion in seawater [10–18]. The biofilm that forms on any material
immersed in seawater is intrinsically heterogeneous and leads to the coexistence of aerated and
deaerated zones, creating aeration cells and favoring the growth and activity of harmful microorganisms.
Electroactive bacteria may even uptake electrons from the metal and thus influence directly the
corrosion rate [23–25]. However, if sulfate-reducing bacteria (SRB) develop initially in deaerated zones
because they are anaerobe microorganisms and favor temporarily localized corrosion processes, they
rapidly colonize the whole metal surface [26], which ultimately leads again to more or less uniform
corrosion [22,26].
The solid corrosion products forming on the steel surface are the primary consequences of the
dissolution of the metal. They can strongly affect the ongoing corrosion process. First of all, they form
a physical barrier between metal and environment and therefore contribute to the protection of the
metal, because they hinder the transport of dissolved oxygen from seawater to the metal surface [20,21].
Secondly, they are porous and thus offer a unique habitat for microorganisms that can develop in a
specific environment [26–31]. Thirdly, some phases are electronic conductors, e.g., magnetite [32–34]
and iron sulfides [16,23], which can favor galvanic cells. The composition of the corrosion product
layer also varies with the exposure zone [26,35–42] and may change in the long term [43].
This article reports recent advances in the understanding of localized corrosion processes of carbon
steel permanently immersed in natural seawater, obtained through a detailed analysis of the corrosion
product layers [26,30,36,41,42,44]. These layers, formed on steel surfaces in laboratory experiments or
seaport exposure sites, reflect the complexity of iron chemistry in natural seawater. Their composition
differs in anodic and cathodic zones [9,41,42] so that they participate actively in the persistence of
corrosion cells and favor localized corrosion processes [41].

2. Materials and Methods

2.1. Materials and Exposure Sites


The article is focused on results obtained with natural seawater. Actually, the complex mechanisms
involved during marine corrosion cannot be fully reproduced via laboratory experiments using
artificial seawater.
The results discussed here were obtained with coupons immersed in three main exposure sites.
Two of these sites were set in seaports of La Rochelle (Atlantic Ocean, France). In one site (Les Minimes
marina), the coupons were immersed at constant immersion depth (~50 cm). In the other site (cargo
port), the immersion depth, i.e., the distance between coupon and atmosphere, varied with the tide,
between a few centimeters and ~5 m. The third site was the Naval Group Research laboratory in
Cherbourg (English Channel, France). In this facility, natural seawater is pumped directly from the
English Channel to flow continuously at a rate of 100 L/h through the electrochemical cells. In this last
case, the steel coupons were 5–10 cm distant from the seawater/air interface so that differential aeration
phenomena could be induced and favored localized corrosion. However, the results obtained in each
site proved comparable so that general trends could be deduced. More detailed information on the
exposure conditions can be found in the original articles [26,30,36,37,40–44].
It is also worthy to recall the average composition of seawater. The study of the water flowing
through the Naval Group Research lab in Cherbourg led to the following composition (main ionic
Corros. Mater. Degrad. 2020, 1 200

species only): [Cl− ] = 19 g kg−1 , [SO4 2− ] = 2.7 g kg−1 , [HCO3 − ] = 0.14 g kg−1 , [Na+ ] = 10 g kg−1 ,
[Mg2+ ] = 1.3 g kg−1 , [Ca2+ ] = 0.41 g kg−1 and [K+ ] = 0.4 g kg−1 . It clearly shows that seawater is not,
in any case, a simple NaCl solution, and the role of various other species, e.g., SO4 2− , Ca2+ , Mg2+ ,
and HCO3 − , is clearly illustrated by the composition of the mineral layers covering carbon steel in
seawater [26,30,36,41,42]. Note also that the alkalinity of seawater (pH measured between 8.0 and 8.2
at Naval Group Research lab), is mainly associated with the presence of hydrogencarbonate ions.
The present article describes the behavior of carbon steel and thus does not deal with the influence
of alloying elements. More precisely, the article relates to studies performed with three main materials.
The first one is the alloy typical of sea harbor steel sheet piles, i.e., carbon steel S355GP. The other
alloys are S355NL and TU250B carbon steels. The composition of each steel grade is given in Table 1.
The slight observed differences, e.g., less carbon and manganese and more copper for TU250B, did not
lead to detectable effects [41] so that it can be considered that the composition of the corrosion product
layer does not depend, or only marginally, on the considered carbon steel grade. This is also illustrated
by the results obtained with different carbon steel alloys [26,30,36,41,42]. Experimental details about
the preparation of the carbon steel samples can be found in the original articles [26,30,36,37,40–44].
In any case, the steel surface was shot blasted to remove any trace of previous rusting or mill scale and
degreased with acetone.

Table 1. Composition (wt. %) of the three carbon steels used in this study (nominal composition for
S355GP or as given by the manufacturer, i.e., Arcelor-Mittal for S355NL, and Vallourec for TU250B),
rest = Fe.

Carbon Steel C Mn P S Si Al Cr Cu Ni
S355GP ≤0.27 ≤1.7 ≤0.055 ≤0.055 ≤0.6 0.02 - - -
S355NL 0.17 1.4 0.015 0.005 0.21 0.02 0.02 0.01 0.02
TU250B 0.12 0.6 0.017 0.006 0.22 - 0.08 0.19 0.09

2.2. Characterization of the Corrosion Product Layers


The characterization of the corrosion product layers requires that samples are preserved from
oxidation by air (or more exactly oxygen). This important point is developed in the article (Section 5).
In the lab, the samples can be kept in a freezer (−24b ◦ C), a simple procedure that hinders the
oxidation of Fe(II)-based corrosion products that are reactive towards oxygen [40–42]. This frozen state
moreover facilitates the sampling of small slices of corrosion product layers that can be easily handled.
Once warmed up, these solid slices transform to sludge due to their high content of water.
The thorough and unambiguous characterization of the phases present in the corrosion product
layers cannot be achieved using a single analytical tool. Moreover, the characterization techniques
must be adapted to the particular nature of the layers that are thick, heterogeneous, and very porous.
The pores are filled with water, microorganisms, and organic matter. When analyzing a corrosion
product layer, the first objective is to identify all the crystalline solid phases that constitute this layer.
This can be achieved using X-ray diffraction (XRD). Another technique must at least be used to
characterize amorphous, nanocrystalline or poorly crystallized compounds. The second objective is to
determine how the various compounds are distributed inside the corrosion product layer, in particular,
to establish if there exists some kind of stratification from metal to seawater. A local method such as
µ-Raman spectroscopy (µ-RS) is adequate for that purpose. µ-RS is moreover suitable for the analysis of
wet (and even immersed) samples and the characterization of non-crystalline matter. The combination
of µ-RS with XRD, used in various works [26,30,36,37,39–44], is then a minimal requirement.
µ-RS analysis was performed in any case at room temperature using a Jobin Yvon High Resolution
Raman spectrometer (LabRam HR or LabRam HR evo) equipped with a confocal microscope and a
Peltier-based cooled charge-coupled device (CCD) detector. Two laser sources were used indifferently,
a solid-state diode-pumped green laser (532 nm) or a He-Ne laser (632.8 nm). The laser power must
be reduced down to 10% (0.6 mW), and in some cases, even to 1% (0.06 mW), of the maximum to
Corros. Mater. Degrad. 2020, 1 201

prevent the transformation of the analyzed compounds into hematite α-Fe2 O3 . This transformation
can take place due to excessive heating, as observed for Fe(III)-oxyhydroxides [45] and magnetite
Fe3 O4 [45,46]. The acquisition time was variable and depended on the nature of the analyzed phase.
It was generally equal to 60 s but could be increased up to 2 min to optimize the signal to noise ratio.
At least 20 zones (diameter of 3–6 µm) of a given sample were analyzed through a 50× long working
distance objective (50 × LWD). A motorized XY stage was used in order to focus on the different zones
of the sample and determine the possible organization/stratification of the corrosion product layer.
Most of the investigations could be performed without specific protection from air, which was possible
because the time required for analysis was short and samples remained wet during analysis, which
delays oxidation.
X-ray diffraction (XRD) analysis was performed with an Inel EQUINOX 6000 diffractometer,
using a curved detector (CPS 590), with the Co-Kα radiation (λ = 0.17903 nm) at 40 kV and 40 mA.
The CPS 590 detector is designed for the simultaneous detection of the diffracted photons on a 2θ range
of 90◦ . The analysis was performed with a constant angle of incidence (5 degrees) for 30 to 60 min
depending on the sample. To prevent the oxidation of Fe(II)-based compounds during preparation and
analysis, the samples were mixed with a few drops of glycerol in a mortar before being crushed until a
homogenous oily paste was obtained. With this procedure, the various particles that constitute the
sample are coated with glycerol and thus sheltered from the oxidizing action of O2 [47]. Glycerol may
only give rise to a very broad “hump” visible on the XRD pattern between 2θ~25◦ and 2θ~35◦ [48].
XRD and µ-RS results are displayed throughout the present article to illustrate the complexity of
the corrosion product layers formed on carbon steel in natural seawater.

3. General Features of the Corrosion Product Layer

3.1. Stratification of the Corrosion Product Layer


After several years in natural seawater, the surface of carbon steel structures permanently
immersed in natural seawater is covered by a corrosion product layer itself covered by biofouling.
Figure 1 displays the main features of the layers formed on carbon steel coupons after 6 to 11 years of
immersion [36,41,42,44].

Figure 1. Schematic representation (cross-section) of the layer covering a carbon steel coupon after
6–11 years of permanent immersion in natural seawater.

The corrosion product layer can be several millimeters thick, reaching locally a thickness of
1 cm or more. Under the biofouling layer that may incorporate mineral compounds (sand, clay,
fragments of shells) and macro-organisms, an orange-brown corrosion product layer is generally seen.
This is the outer stratum, mainly composed of Fe(III)-oxyhydroxides (FeOOH). Lying underneath
is found the inner stratum, characterized by a black color, that is in contact with the steel surface.
This inner stratum is mainly composed of Fe(II)-based corrosion products, mixed with magnetite
Fe3 O4 . The Fe(II)-based compounds are reactive towards oxygen and the Fe(III)-oxyhydroxides that
constitute the orange-brown outer layer are the end products of their oxidation by dissolved O2 . After
6–11 years in natural seawater, the inner black stratum is much thicker than the orange-brown outer
stratum. This demonstrates that the steel surface and the main part of the corrosion product layer
Corros. Mater. Degrad. 2020, 1 202

are no more reached by dissolved oxygen, i.e., that anoxic conditions are met. This explains why
anaerobe microorganisms develop systematically inside the corrosion product layers and influence
necessarily the corrosion process, as perfectly illustrated by the phenomenological model proposed by
Melchers et al. [17,22]. Dissolved oxygen cannot reach the inner part of the corrosion product layer
because the aerobe microorganisms that colonize the biofouling layer and the orange-brown outer
stratum consume it. The little oxygen that could possibly reach the dark inner stratum would then
react with the Fe(II)-based corrosion products to form FeOOH phases and/or magnetite.
The stratification of the corrosion product layer can be clearly observed as soon as 1–2 months of
immersion [30].

3.2. Composition of the Corrosion Product Layer


Figure 2 shows the XRD pattern of the corrosion product layer that covered a carbon steel coupon
immersed 6 months in natural seawater (Les Minimes marina exposure site, La Rochelle, Atlantic
Ocean). The main crystalline solid phases that constitute such corrosion product layers are all detected
here. Two Fe(III)-oxyhydroxides, namely goethite α-FeOOH and lepidocrocite γ-FeOOH, are identified.
They constitute the orange-brown outer stratum. Magnetite, the Fe(II,III) mixed-valence oxide Fe3 O4 ,
is identified too. It is present in the inner dark layer. Magnetite can form in anoxic conditions as shown
by thermodynamic data e.g., [49,50]. The overall reaction involves the oxidation of three Fe(0) atoms to
one Fe(II) cation and two Fe(III) cations, and the reduction of four water molecules to hydrogen. It can
be written as:
3Fe + 4H2 O → Fe3 O4 + 4H2 (1)

This first process explains the formation of magnetite on or close to the metal surface. However,
magnetite can also be obtained indirectly as an oxidation product of Fe(II)-based corrosion products at
very low oxygen flow rates [51]. This second process explains why magnetite was in some cases mainly
detected at the interface between the inner dark stratum and the outer orange stratum [41], where a
small amount of oxygen can access after having crossed over the outer parts of the biofouling/corrosion
product layer.

Figure 2. XRD analysis of the corrosion product layer covering a carbon steel coupon after
6 months in natural seawater (exposure site: Les Minimes marina, La Rochelle, Atlantic Ocean).
A = aragonite, G = goethite, GR = sulfate green rust, GC = carbonate green rust, GCl = chloride green
rust, L = lepidocrocite, M = magnetite, and Q = quartz. The diffraction lines are denoted with the
corresponding Miller index.
Corros. Mater. Degrad. 2020, 1 203

The other corrosion products identified via the XRD pattern of Figure 2 are green rust (GR)
compounds, namely the sulfate GR, the carbonate GR, and the chloride GR. Green rust compounds
are mixed-valence Fe(II,III) layered double hydroxides (LDH) mainly containing Fe(II) cations (67%
to 75%). In contrast, magnetite contains mostly Fe(III) cations (67% i.e., only 33% of Fe(II) cations).
The crystal structure of GR compounds consists of the stacking of hydroxide layers, built on Fe(OH)6
octahedra, carrying a positive charge because of the presence of the Fe(III) cations [52–55]. So-called
interlayers constituted of anions and water molecules are intercalated between two adjacent hydroxide
layers. Various GRs can be obtained depending on the anions present in the surrounding environment,
and seawater, given its composition, can induce the formation of GR(Cl− ), GR(SO4 2− ), and GR(CO3 2− ).
The sulfate green rust is predominant as revealed by the numerous results accumulated in the
past ten years [26,30,36,41,42]. This is also illustrated in Figure 2 by the higher relative intensity of
the diffraction peaks of GR(SO4 2− ). The chloride GR, which could be expected to be the predominant
GR variety because of the high chloride concentration of seawater, was rarely identified or as a very
minor compound. However, the stability of the various GR compounds does not only depend on
the intercalated anion concentration. It depends on the geometry, size, and number of charges of the
anion, or more generally, on any parameter that influences the stability of the LDH crystal structure.
The studies dealing with the affinity of this structure for various anions revealed that divalent anions
were generally preferred to monovalent ones [56–58]. In solutions characterized by large [Cl− ]/[SO4 2− ]
concentration ratios, it was demonstrated accordingly that GR(SO4 2− ) was obtained preferentially to
GR(Cl− ) [58].
Other compounds can be identified with the XRD pattern of Figure 2. Quartz, i.e., sand, comes
from the environment. Aragonite, a CaCO3 phase, can also come from the environment as some shells
are made of it. However, its formation may be associated with the corrosion process, as explained in
Section 4.1.
Table 2 lists the compounds positively identified via XRD and/or µ-RS spectroscopy in the corrosion
product layers formed on carbon steel permanently immersed in natural seawater, as reported in
references [26,30,36,41,42,44]. Chukanovite Fe2 (OH)2 CO3 and greigite Fe3 S4 were rarely identified and
do not appear among the products revealed by the XRD pattern of Figure 2. In contrast, mackinawite
FeS is one of the main corrosion products, present in the inner dark layer, but it is not detected either via
the XRD pattern. This is generally the case because mackinawite is mainly found in a nanocrystalline
state and is rarely identified by XRD [26,30,36,41,42,44]. It can however be detected using µ-Raman
spectroscopy [59].

Table 2. List of corrosion products/minerals identified on carbon steel coupons or structures permanently
immersed in natural seawater according to references [26,30,36,41,42,44] and their possible link with
the anodic/cathodic nature of the underlying metal surface.

Compound Name: Chemical Link to Anodic and Cathodic


Importance/Abundance and Localization
Formula Zones
Sulfate green rust, GR(SO4 2− ):
Main component of the inner layer Favored in anodic zones
FeII 4 FeIII 2 (OH)12 SO4 ·8H2 O
Carbonate green rust, GR(CO3 2− ):
Minor component of the inner layer Favored in cathodic zones
FeII 4 FeIII 2 (OH)12 CO3 ·2H2 O
Chloride green rust, GR(Cl− ):
Minor component of the inner layer Not known
FeII 3 FeIII (OH)8 Cl·2H2 O
Chukanovite: Fe2 (OH)2 CO3 Rare Cathodic zones
Mackinawite/Fe(III)-containing
Main component of the inner layer, due to SRB Not known
mackinawite: FeII S/ FeII 1-3x FeIII 2x S
Greigite: Fe3 S4 Rare Not known
Magnetite: Fe3 O4 Main component of the inner layer Favored in cathodic zones
Lepidocrocite: γ-FeOOH Main component of the outer layer Anodic zones or uniform corrosion
Goethite: α-FeOOH Main component of the outer layer Anodic zones or uniform corrosion
Akaganeite: β-FeOOH Rare, minor component Not known
Aragonite: CaCO3 Outer layer Cathodic zones
Corros. Mater. Degrad. 2020, 1 204

Figure 3 shows a picture obtained with the optical microscope of the µ-RS apparatus. It was taken
during the analysis of the inner dark layer covering a carbon steel coupon immersed 6 years in natural
seawater (exposure site: Naval Group laboratory, Cherbourg, English Channel). At the center of the
image, a stack of thin bluish platelets can be seen. Only a part of these platelets is visible, as the rest is
covered with shiny black matter, but one can see that these particles seem to have the characteristic
hexagonal shape of GR crystals [36,41,60–62]. µ-RS analysis of these platelets (not shown) confirmed
that they corresponded to a GR compound. The shiny black matter that surrounds them was also
analyzed by µ-RS (Figure 4). Figure 4a displays the most commonly found spectrum. It is composed
of two Raman peaks, the main one located at 283 cm−1 and the other one at 207 cm−1 . This spectral
signature is typical of nanocrystalline mackinawite [59]. This nanocrystalline compound, formerly
considered as “amorphous FeS” or “poorly ordered FeS”, is the solid that precipitates from dissolved
Fe(II) and S(-II) species [63,64]. Aging in the solution can improve the crystallinity but this process is
very slow in alkaline conditions and at ambient temperatures [59]. In seawater (pH~8.1), mackinawite
then remains essentially in its nanocrystalline state.

Figure 3. Photograph of a stack of hexagonal platelets of GR(SO4 2− ) surrounded by shiny black matter
(Optical microscope of the Raman apparatus, ×50 objective), taken during the analysis of the inner
stratum of the corrosion product layer covering cathodic zones of a carbon steel coupon after 6 years in
natural seawater (exposure site: Naval Group laboratory, Cherbourg, English Channel).

Figure 4. Raman spectra obtained from the shiny black matter shown in Figure 3: (a) nanocrystalline
mackinawite, (b) mixture of various iron sulfides.
Corros. Mater. Degrad. 2020, 1 205

Figure 4b displays another Raman spectrum of the shiny black matter, less frequently observed
than the previous one. First, the two Raman peaks of nanocrystalline mackinawite are seen, at 207 cm−1
and 282 cm−1 . Secondly, other intense peaks are seen at 125 cm−1 , 253 cm−1 , 310 cm−1 , and 364 cm−1 .
They are attributed to the so-called “Fe(III)-containing mackinawite”, a slightly oxidized form of
mackinawite [59,65] that may contain up to 20% of Fe(III) [59,66]. The proposed chemical formula
is then FeII 1-3x FeIII 2x S [59]. The in-situ oxidation of Fe(II) into Fe(III) can take place even in anoxic
conditions, and this process finally leads to greigite Fe3 S4 [65] via a solid-state transformation [67].
Actually, the spectrum of Figure 4b shows typical features of greigite that are the peaks at 143 cm−1 ,
194 cm−1 , and 364 cm−1 (shared with Fe(III)-containing mackinawite) [65,68].
The iron sulfides are formed because of bacterial activity. Sulfide species are not present in
seawater except in deaerated conditions that promote the growth and activity of sulfide-producing
bacteria such as SRB. In studies combining a microbiological analysis of the bacterial flora with a
characterization of the corrosion products, the presence of FeS was indeed associated with that of
SRB (e.g., [18,26,27,30,69]). As explained above, anoxic conditions are established after some time in
the inner black stratum of the corrosion product layers and consequently anaerobe microorganisms
such as SRB can grow and be active. At the beginning of the corrosion process, only a few deaerated
regions are found and FeS forms only locally. It could be detected in one particular zone of a steel
coupon after 1 month of immersion, a finding associated with the concomitant detection of SRB [30].
After 6–12 months, FeS is scattered more or less homogeneously in the inner black layer [26], which
indicates that anaerobic conditions prevail all over the steel surface. This corresponds to the moment
when the influence of SRB becomes significant, as it affects the whole steel surface, a particular time of
the corrosion process that is characterized by an increase of the corrosion rate [22,69].
Finally, maghémite (γ-Fe2 O3 ) and ferrihydrite (Fe5 HO8 ·4H2 O) are two compounds possibly
present as minor components of the orange outer stratum. They are difficult to characterize by XRD
because: (i) the XRD pattern of maghémite is close to that of magnetite and (ii) ferrihydrite is a very
poorly ordered and crystallized form of Fe(III)-(oxy)hydroxide. They could not be unambiguously
identified in previous works [26,30,36,41,42,44] and for this reason, were omitted from Table 2.

3.3. GR(SO4 2− ), the First Compound Resulting of the Corrosion of Carbon Steel in Seawater
Analysis of the corrosion product layers formed on carbon steel coupons after 1 week of immersion
in natural seawater (Les Minimes marina, La Rochelle, Atlantic Ocean) revealed only two phases,
the sulfate GR and lepidocrocite γ-FeOOH [30]. Actually, lepidocrocite is the main oxidation product
of GR(SO4 2− ) by dissolved O2 [58], which indicates that GR(SO4 2− ) is the first solid phase to form from
the dissolved species produced by the corrosion of the metal. This point was confirmed by anodic
polarization experiments performed with carbon steel electrodes immersed in artificial seawater-like
solutions deaerated with argon [36]. After 15 min, 3 h, and 26 h of polarization, only the sulfate GR
(with traces of the carbonate GR) could be identified on the steel surface.
The formation of GR compounds in seawater-like conditions was also studied using another
method. GR compounds could be precipitated by mixing a solution of NaCl, Na2SO4·10H2O, FeCl2·4H2O,
and FeCl3 ·6H2 O with a solution of NaOH [70]. The obtained precipitates were analyzed immediately
after their formation and after 1 week of aging in suspension at room temperature. The initial precipitate
was in any case a mixture of GR(SO4 2− ) with GR(Cl− ). After one week of aging, only traces of GR(Cl− )
remained, GR(SO4 2− ) being the final result of the overall process. This study shows that GR(Cl− ) may
be an intermediate transient compound, rapidly transformed to GR(SO4 2− ). This would explain why
GR(Cl− ) could be sometimes identified, always in small amounts, in the dark inner stratum of the
corrosion product layer.
From all these findings, it can be proposed that GR(SO4 2− ) forms from the dissolved Fe(II) species
produced by the corrosion of steel through the process illustrated in Figure 5. This process may take
place within a few minutes or a few seconds as GR(SO4 2- ) was the only product detected after 15 min
of immersion [36].
Corros. Mater. Degrad. 2020, 1 206

Figure 5. Schematic representation (a–c) of the possible mechanism for GR(SO4 2− ) formation as the
first solid phase resulting from the corrosion of carbon steel in seawater. The water molecules present
in the interlayer of the GR or adsorbed on the hydroxide sheets are omitted for clarity.

It is generally admitted that the Fe2+ cations produced by the anodic reaction precipitate first with
the OH− ions produced by the cathodic reaction (see Section 4.1), which leads to the Fe(II) hydroxide
Fe(OH)2 . The crystal structure of Fe(OH)2 is based on sheets formed by hydroxide-bridged FeII (OH)6
octahedra. The formation of the solid phase may then begin by the formation of small units of such
sheets, where Cl− and SO4 2− anions (predominant in seawater) are adsorbed (Figure 5a). At the pH of
seawater, Fe(OH)2 can be oxidized to GR(SO4 2− ) even in anoxic conditions [50], so that in any case
some Fe(II) cations are rapidly oxidized to Fe(III). A FeIII (OH)6 octahedron necessarily bears an excess
positive charge and attracts anions, facilitating their adsorption on the forming hydroxide sheets,
as illustrated in Figure 5b with the example of a sulfate anion. The stacking of such Fe(III)-containing
hydroxide sheets, with intercalated anions and water molecules, actually corresponds to the structure
of GR compounds [52–55], as explained in previous Section 3.1. Since divalent anions give greater
stability to this structure, Cl− ions should be finally expelled from the interlayers, finally leading to
GR(SO4 2− ) as illustrated in Figure 5c.
The overall process leading from the Fe(II) dissolved species to the first solid corrosion product,
i.e., GR(SO4 2− ), can then be summarized by the reactions (2), if dissolved O2 is present, or (3), if anoxic
conditions are established:
1
6Fe2+ + 10OH− + SO4 2− + O2 + 9H2 O → FeII 4 FeIII 2 (OH)12 SO4 ·8H2 O (2)
2

6Fe2+ + 10OH− + SO4 2− + 10H2 O → FeII 4 FeIII 2 (OH)12 SO4 ·8H2 O + H2 (3)

4. Characterization of Anodic and Cathodic Zones

4.1. Heterogeneity of the Corrosion Product Layer and Role of Interfacial pH


The use of sufficiently large carbon steel coupons, e.g., 10 cm × 10 cm sized coupons, generally
enables us to observe the heterogeneity of the corrosion process [41]. An example is displayed in
Corros. Mater. Degrad. 2020, 1 207

Figure 6, showing the surface of a coupon after 6 years in seawater. Although the coupon surface
is mostly covered by biofouling and macro-organisms (a shell is visible near the center) different
zones can be easily distinguished. At the bottom (middle of the lower edge) of the coupon, a large
orange tubercle has formed. This accumulation of corrosion products indicates that the corrosion
process has been accelerated here. The underlying metal surface then more likely corresponds to an
anodic zone. This can be rigorously demonstrated after the corrosion product layer is removed via the
determination of the localized degradation depth [9,41,42]. Other zones are covered by a much thinner
dark layer of corrosion products, as indicated in Figure 6, where the corrosion rate was less important.
They correspond to cathodic zones of the underlying carbon steel surface.

Figure 6. Picture showing the surface of a 10 cm × 10 cm carbon steel coupon after 6 years in natural
seawater (exposure site: Naval Group laboratory, Cherbourg, English Channel).

The formation of cathodic zones, where the cathodic reaction rate is higher than that of the anodic
reaction rate, and anodic zones, where the cathodic reaction rate is lower than the anodic reaction
rate, may have various origins. The possible causes of localized corrosion are well established and
include the inherent heterogeneity of the biofilm, the differential aeration cells (especially for vertical
surfaces), and the heterogeneity of the steel surface (inclusions, welds, segregation . . . ). The possible
persistence of the galvanic coupling between anodic and cathodic zone is however governed, for long
immersion times, by the properties of the corrosion product layers that progressively grow on each
zone. The detailed characterization of the products forming in cathodic and anodic zones was then a
key point towards the understanding of the long-term localized corrosion processes of carbon steel in
seawater [41,42].
The thick tubercles of corrosion products such as the one shown in Figure 6 are, like the corrosion
product layers resulting from uniform corrosion, composed of two main strata, an inner dark stratum
in contact with the metal surface and an outer orange-brown stratum [41,42]. Figure 7 displays the
XRD pattern of a fragment of the inner dark stratum sampled close to the metal surface in an anodic
zone. This pattern is mainly composed of the diffraction peaks of the sulfate GR that are very intense.
Other GR compounds are not identified and only the main diffraction peak of magnetite (M311) is
clearly seen.
Corros. Mater. Degrad. 2020, 1 208

Figure 7. XRD analysis of the inner stratum of the corrosion product layer covering anodic zones
of a carbon steel coupon after 6 years in natural seawater (exposure site: Naval Group laboratory,
Cherbourg, English Channel). GR = sulfate green rust, and M = magnetite. The diffraction lines are
denoted with the corresponding Miller index.

Figure 8 displays the XRD pattern of a fragment of the whole dark layer that covered a cathodic
zone of the metal surface. The most intense peak is in this case the main diffraction peak M311 of
magnetite. The diffraction peaks of other GR compounds are now clearly detected, and those of the
carbonate GR are particularly intense. Aragonite is also clearly identified. The Fe(III)-oxyhydroxides
are minor components of the corrosion product layer formed in this cathodic area, in agreement with
its visual aspect (black color). This demonstrates that the composition of the corrosion product layer
depends on the anodic/cathodic character of the underlying metal surface.

Figure 8. XRD analysis of the inner stratum of the corrosion product layer covering cathodic zones
of a carbon steel coupon after 1 year in natural seawater (exposure site: cargo port of La Rochelle,
Atlantic Ocean). A = aragonite, GR = sulfate green rust, GC = carbonate green rust, GCl = chloride
green rust, L = lepidocrocite, M = magnetite, and Q = quartz. The diffraction lines are denoted with the
corresponding Miller index.
Corros. Mater. Degrad. 2020, 1 209

Iron sulfides were identified in any case in both anodic and cathodic zones using µ-RS. Actually,
the Raman spectra of mackinawite displayed in Figure 4 were obtained from the corrosion product
layer covering a cathodic zone. This demonstrates that SRB can grow and be active in both kinds of
zones, which mainly indicates that oxygen does not reach the inner dark layer neither in the anodic
nor in the cathodic zones (this point is further discussed in Section 4.4). At the present time, it is not
possible to state whether SRB develops first or preferentially in one kind of zone.
In Table 2 that lists the products positively identified inside the corrosion product layers [26,30,36,41,42,44],
it is indicated whether a compound is preferentially formed in anodic or cathodic zones. The compounds
favored in cathodic zones are carbonate-containing phases, namely aragonite, chukanovite, and
GR(CO3 2− ), and magnetite. In each case, the main reason is the increase of the interfacial pH induced
by the difference between the cathodic and the anodic reaction rates. On a metal surface where uniform
corrosion takes place, the cathodic and anodic reaction rates are equal. The anodic reaction produces
Fe2+ cations:
Fe → Fe2+ + 2e− (4)

The corresponding cathodic reaction, whether it is O2 or H2 O reduction, produces 2 OH− ions per
2 consumed electrons:
1
O + H2 O + 2e− → 2OH− (5)
2 2
2H2 O + 2e− → H2 + 2OH− (6)

Fe2+ and OH− ions are then produced in a ratio OH− /Fe2+ = 2 and subsequently incorporated
inside the solid phases that constitute the corrosion product layer. This is commonly summarized by
the following reaction:
Fe2+ + 2OH− → Fe(OH)2 (7)

As discussed in Section 3.3, the formation of Fe(II)-hydroxide sheets is rapidly followed in seawater
by that of GR(SO4 2− ) but, as shown by reaction (7), the pH at the steel/seawater interface is unaffected.
The same conclusion is drawn if another corrosion product is considered, as illustrated by the example
of magnetite formation (reaction 1). In cathodic zones, the rates of reduction reactions (5) and/or (6)
is higher than that of the anodic reaction (4) because part of the consumed electrons comes from the
anodic zones. The ratio between the produced OH− and Fe2+ ions in a cathodic zone is then higher
than 2, which induces an increase of the interfacial pH. This effect is well-known for steel structures
subjected to cathodic protection and leads to the formation of the so-called calcareous deposit, mainly
composed of aragonite CaCO3 e.g., [44,71–74]. The presence of aragonite among the corrosion products
of steel inside the layers forming on a cathodic zone is then clearly associated with the increase of
the interfacial pH. This assumption was confirmed via experiments performed in artificial seawater
where shells, the other possible origin of aragonite, were of course not present [9]. The increase of the
interfacial pH leads to changes in the inorganic carbonic equilibrium at the metallic interface and favor
CO3 2− over HCO3 − :
HCO3 − + OH− → CO3 2− + H2 O (8)

This explains why the formation of aragonite is not the only process favored in the cathodic zones:
the formation of other CO3 2− based compounds, i.e., chukanovite and GR(CO3 2− ), is also facilitated.
In contrast, the ratio between the produced OH− and Fe2+ ions is lower than 2 in the anodic zones.
As Fe2+ is a Lewis acid, the interfacial pH may tend to decrease in the anodic areas. This well-known
phenomenon is the origin of various localized corrosion phenomena such as crevice corrosion, pitting
corrosion, etc...

4.2. Equilibrium Conditions between GR(SO4 2− ) and GR(CO3 2− )


The formation of GR(CO3 2− ) in the cathodic zones was thoroughly studied. The first results were
obtained with a study of the impact of cathodic protection on aged (5 years of immersion) and already
Corros. Mater. Degrad. 2020, 1 210

strongly corroded carbon steel coupons. It was observed after one year of cathodic polarization that
the whole amount of GR(SO4 2− ) initially present in the corrosion product layers had been entirely
transformed to GR(CO3 2− ) [44]. The reaction induced by the polarization can be written as follows:

FeII 4 FeIII 2 (OH)12 SO4 ·8H2 O + CO3 2− → FeII 4 FeIII 2 (OH)12 CO3 ·2H2 O + 6H2 O + SO4 2− (9)

This writing clearly shows that the transformation is favored by the increase of the CO3 2−
concentration. More precisely, the equilibrium conditions between GR(SO4 2− ) and GR(CO3 2− ) are
governed by the pH, the sulfate concentration, and the carbonate species concentration [9,44]. This can
be highlighted by writing the reaction with the main carbonate dissolved species present in seawater,
i.e., HCO3 − :

FeII 4 FeIII 2 (OH)12 SO4 ·8H2 O + HCO3 −  FeII 4 FeIII 2 (OH)12 CO3 ·2H2 O + 6H2 O + SO4 2− + H+ (10)

The equilibrium conditions are then expressed by:

pK = pH – log[a(SO4 2− )/a(HCO3 − )] (11)

In Equation (11), a(SO4 2− ) and a(HCO3 − ) are the activities of SO4 2− and HCO3 − in the solution
where GR(SO4 2− ) and GR(CO3 2− ) coexist. A study of these equilibrium conditions was achieved to
obtain a numerical value of the equilibrium constant [9], which led to:

pK = 7.85 ± 0.35 (12)

Using the typical values of a(SO4 2− ) and a(HCO3 − ) for seawater [9], it was computed that the pH
corresponding to the equilibrium conditions between both GR compounds was equal to 8.24 ± 0.35 [9],
which is similar or only slightly higher than the average pH of seawater. This explains why a slight
increase of the interfacial pH can indeed favor the formation of GR(CO3 2− ) at the detriment of that
of GR(SO4 2− ).

4.3. Variations of the Magnetite Content


The key point is however the competitive formation processes of magnetite and GR(SO4 2− )
because magnetite is a well-known electronic conductor [32–34] whereas, to our better knowledge,
GR(SO4 2− ) is an insulator. According to thermodynamic data, magnetite would indeed be favored by
an increase of pH, and conversely, GR(SO4 2− ) by a decrease of pH [50]. This was also confirmed by
experimental results [50].
As for the corrosion product layers formed on carbon steel permanently immersed in seawater,
the relative proportions of magnetite and GR(SO4 2− ) proved to depend on the anodic/cathodic character
of the underlying metal surface. This is illustrated by the results synthesized in Table 3.
This table gathers data from [41,42] and unpublished results coming from the same research
program. To quantify the changes in the relative proportions of magnetite and GR(SO4 2− ) revealed
by XRD analysis, the M311 diffraction line intensity to GR003 diffraction line intensity ratio, i.e.,
I(M311 )/I(GR003 ), was used. This ratio was chosen because, in the considered experimental conditions
(Naval Group laboratory, Cherbourg, English Channel), it was close to 1 (actually between 0.77 and
1.05) for the corrosion product layers generated by uniform corrosion [41]. After 6 years of immersion
in these conditions, the average corrosion rate was determined, from thickness loss measurements,
to be equal to 0.07 ± 0.02 mm/year for uniform corrosion. Therefore, a zone of the metal surface where
the local corrosion rate is significantly higher than 0.07 mm/year should be considered as an anodic
zone. Conversely, if the local corrosion rate is significantly lower than 0.07 mm/year then this part of
the steel surface should be considered as a cathodic zone.
Corros. Mater. Degrad. 2020, 1 211

Table 3. Determined values of the I(M311)/I(GR003) ratio for inner layers on anodic and cathodic zones
and associated corrosion rates from References [41,42] and previously unpublished results from the
same research program. The corrosion rates were determined by local thickness measurements as
described in [40–42] on carbon steel coupons permanently immersed 6–8 years in natural seawater
(exposure site: Naval Group laboratory, Cherbourg, English Channel).

Zone Corrosion Rate (mm/Year) I(M311)/I(GR003)


Strongly anodic 0.50 ± 0.01 0
Anodic 0.20 ± 0.03 0.12
Anodic 0.15 ± 0.04 0.5
Anodic 0.14 ± 0.02 0.03
Slightly anodic 0.08 ± 0.02 0.51
Uniform corrosion 0.07 ± 0.02 0.77 to 1.05
Slightly cathodic 0.06 ± 0.01 0.5
Slightly cathodic 0.06 ± 0.01 3.1
Cathodic 0.05 ± 0.01 1.9 to 3.3
Cathodic 0.040 ± 0.005 8.9 to 10
Cathodic 0.035 ± 0.005 2.6
Strongly cathodic ~ 0.01 4.9 to 14.3

The data related to the various analyzed zones are listed in Table 3 with decreasing measured local
corrosion rates or, in other words, from the most severely degraded areas (0.5 mm/year, top line) to
less degraded regions (0.01 mm/year, bottom line). In each case, the corrosion product layer covering
the metal was analyzed by XRD so that the composition of the layer could be associated with the
corrosion rate. It can then be seen in Table 3 that the I(M311 )/I(GR003 ) ratio determined from the XRD
analysis increases with decreasing corrosion rate. This trend is more clearly illustrated by the diagram
displayed in Figure 9, where the average value of the I(M311 )/I(GR003 ) ratio is plotted against the
average local corrosion rate. This clearly shows that the magnetite content of the corrosion product
layer increases with the cathodic character of the metal surface.

Figure 9. Variation of the I(M311)/I(GR003) ratio (determined from XRD analysis of the dark inner
layers covering steel surfaces after 6 years of immersion in natural seawater), with the corresponding
average local corrosion rate (determined by local thickness measurements as described in [40–42]).

4.4. The Localized Corrosion Mechanism Associated with Magnetite-Rich Cathodic Zones
As described in the previous section, the corrosion product layers covering cathodic zones are
significantly enriched with magnetite that is an electronic conductor. Moreover, these layers do not
Corros. Mater. Degrad. 2020, 1 212

contain, or only in small amounts, FeOOH compounds [41,42], in agreement with visual observations:
these layers are black, i.e., the orange-brown outer stratum composed of FeOOH compounds did not
form. This is only possible if dissolved O2 does not react with the Fe(II)-based compounds present
in the corrosion product layer. This means that the O2 molecules reaching the outer surface of the
corrosion product layer are consuming, through the reduction process, electrons coming from another
part of the system, i.e., necessarily the anodic zones of the steel surface. This process is possible only if
an electronic pathway exists inside the corrosion product layer. This pathway can be constituted by a
network of interconnected magnetite particles also connected to the metal surface. Such a network may
indeed be present in the magnetite-rich layers covering the cathodic zones, that are moreover much
thinner than those covering the anodic zones ([41,42], see also Figure 6). A schematic representation of
the mechanism is presented in Figure 10.

Figure 10. Schematic representation of the localized corrosion process associated with the heterogeneity
of the corrosion product layer.

Actually, the anodic zones of the metal surface are thick and enriched with insulating compounds
such as Fe(III)-oxyhydroxides and GR(SO4 2− ). Dissolved O2 cannot reach the metal surface underneath
or be reduced at the surface of magnetite particles that are present in small amounts and consequently
not interconnected. It is consumed in the outer stratum by aerobe microorganisms or in the dark inner
stratum through its reaction with Fe(II)-based corrosion products.
Localized corrosion can only lead to severe issues if it can persist for a long time. The particular
mechanism associated with the heterogeneity of the corrosion product layers is based on the assumption
that dissolved O2 participates in the corrosion process because it can be reduced on the outer surface of
the magnetite-rich layer covering the cathodic zones [41]. Since the formation of magnetite is favored
by an increase of pH, this mechanism has then a self-sustaining ability as it creates conditions that
favor its own persistence: If oxygen reduction takes place preferentially in the cathodic zones, then
the interfacial pH increases and consequently the formation of magnetite is favored. Conversely,
the process favors a decrease of pH in the anodic zones that hinders the formation of magnetite and
leads to an accumulation of insulating compounds on the metal surface.
The role of SRB in this mechanism may also be important. Anaerobic conditions are met at
the metal surface in both anodic and cathodic zones and SRB can grow and be active in both zones,
as demonstrated by the identification of FeS in any case. In the cathodic zones, dissolved O2 may
reach the outer part of the dark corrosion product layer covering the metal before to be reduced at the
surface of magnetite particles so that SRB would preferentially develop closer to the metal. In any case,
iron sulfides are electronic conductors [16,23] and their formation in the cathodic zones would then
favor the persistence of the corrosion cell described above thus reinforcing the corresponding localized
Corros. Mater. Degrad. 2020, 1 213

corrosion mechanism. Electroactive SRB [23–25] could also take advantage of the interconnected
Fe3 O4 /FeS particles network linked to the metal surface in the cathodic zones, and favor its persistence.
Finally, this process would stop once:

(i) the layer covering the cathodic zone is covered by a thick layer of biofouling/calcareous deposit
that prevents (as in anodic zones) dissolved O2 to reach magnetite particles,
(ii) the corrosion product layer covering the cathodic zone becomes so thick that the network of
interconnected Fe3 O4 /FeS particles is no more linked to the metal surface.

5. Oxidation of the Corrosion Product Layers and FORMATION of Akaganeite


Akaganeite (β-FeOOH) is a common corrosion product of steel in marine atmospheres e.g., [38,39,75,76].
Though it is considered as the β-phase of Fe(III) oxyhydroxides, it necessarily contains Cl− ions and its
chemical formula is more exactly FeO1−x (OH)1+x Clx [77]. However, akaganeite was rarely observed,
and always as a minor component, in the studies [26,30,36,41,42,44] dealing with corrosion products
formed on carbon steel permanently immersed in seawater.
The corrosion processes involved during permanent immersion are of course different from those
of atmospheric corrosion. The wet/dry cycles typical of atmospheric corrosion induce oxidation and
transformation of Fe(II) species/compounds that do not take place during permanent immersion and
may be responsible for the formation of akaganeite. To clarify this point a coupon was, after 1 year
of immersion in natural seawater, exposed 11 days to the atmosphere of the laboratory so that the
corrosion product layer dried completely. This layer was then analyzed by XRD and the result is
shown in Figure 11.

Figure 11. XRD analysis of the inner stratum of the corrosion product layer of a carbon steel coupon after
1 year in natural seawater (exposure site: cargo port of La Rochelle, Atlantic Ocean) and subsequent
exposure of 11 days to the atmosphere (laboratory). Ak = akaganeite, A = aragonite, C = calcite,
G = goethite, H = halite (NaCl), L = lepidocrocite, and M = magnetite. The diffraction lines are denoted
with the corresponding Miller index.

The Fe(III)-based compounds already formed during the immersion period, i.e., magnetite,
lepidocrocite, and goethite, are the main components of the dried oxidized layer. The Fe(II)-based
corrosion products, i.e., GR compounds, are not detected anymore. They were integrally oxidized and
transformed into Fe(III) compounds during the 11 days of exposure to air. Aragonite is also identified,
it comes more likely from a cathodic zone of the metal. Calcite is also present but may come from
fragments of shells. The drying of the sample has moreover induced the formation of NaCl (halite).
Corros. Mater. Degrad. 2020, 1 214

Finally, another Fe(III) compound, not initially present in the corrosion product layer, is now identified:
it is akaganeite.
This shows that akaganeite may mostly result from the oxidation and drying of corrosion product
layers previously formed during the immersion of steel in seawater. Actually, the formation of
akaganeite requires a high concentration of both dissolved Fe(II) and Cl− species [78]. During the
drying of a corrosion product layer, water evaporates progressively and dissolved species (Cl− , Fe2+ )
concentrations increase continuously until the formation of akaganeite is possible. The overall reaction
could be written as:

4Fe2+ + O2 + 6H2 O + 4HCl → 4FeO1−x (OH)1+x Clx + 8H+ (13)

Note that intermediate compounds/species could be involved, e.g., GR(Cl− ) [78,79],


Fe(II)-hydroxychloride [78,79] or dissolved Fe(III) species.
This result points out the necessity to shelter the corrosion product layer from the air before
and during analysis. First, the Fe(II)-based compounds must be preserved because, as they form
directly from the dissolution of steel, they convey the most important information about the corrosion
mechanism. Secondly, the formation of compounds not really associated with the corrosion process
must be avoided because it may lead to a misinterpretation of the mechanisms involved.
A final corrosion product must be discussed, the Fe(II)-hydroxychloride β-Fe2 (OH)3 Cl known
as a precursor of akaganeite [78,79]. Though it has not been clearly identified on “modern” steel
structures, it has however been reported as the main component of the corrosion product layer covering
Gallo-Roman iron ingots that remained at the bottom of the Mediterranean Sea for 2000 years [43].
This finding shows that the corrosion mechanisms may change drastically after a (very) long time of
permanent immersion in seawater. However, no information is currently available about this possible
transition period and the associated mechanisms and processes.

6. Conclusions
The thorough characterization of the corrosion product layers provided important information
that led to a better understanding of the mechanisms of marine corrosion of carbon steel permanently
immersed in natural seawater. Important facts must be pointed out:

- It is generally admitted that, after a sufficiently long immersion time (variable but at least
6 months), anoxic conditions are met at the steel surface and inside the inner part of the corrosion
product layer. Anaerobe microorganisms such as SRB can develop and be active, which leads
to the formation of iron sulfides. Marine corrosion is then intrinsically a biologically influenced
process and its complexity cannot be entirely mimicked by laboratory experiments and/or using
artificial seawater. The biofilm itself and other microorganisms are also known to have an
influence on corrosion processes.
- The first solid phase to form on the steel surface is the sulfate GR. It is favored, with respect
to magnetite and carbonate GR, in the anodic zones, by the decrease of the interfacial pH.
A mechanism is proposed, involving (i) the adsorption of anions (mainly Cl− and SO4 2− ) on the
nuclei of Fe(OH)2 hydroxide sheets, (ii) the oxidation of part of the Fe(II) cations to Fe(III) in
the hydroxide sheets and (iii) the stacking of the sheets leading to GR(SO4 2− ) after the release of
Cl− ions.
- A particular mechanism can however involve oxygen even if anoxic conditions are established
at the steel surface. The pH proved to have a significant influence on the composition of the
corrosion product layer and this composition then depends on the anodic/cathodic character of
the underlying metal surface. In cathodic zones, the increased interfacial pH favors the formation
of magnetite (among other compounds) that is an electronic conductor. Associated with a low
corrosion rate, the process then leads to a magnetite-rich layer that remains moderately thick
(~2–5 mm) even after 6–8 years [41,42]. The reduction of dissolved O2 can then take place at the
Corros. Mater. Degrad. 2020, 1 215

outer surface of this layer as long as the magnetite particles remain interconnected and connected
to the steel surface.
- The activity of SRB (and other sulfide-producing bacteria) in the cathodic zones may favor and
reinforce this mechanism because it generates additional conductive corrosion products, i.e.,
iron sulfides.
- Finally, it is necessary to preserve the samples from air to avoid the transformation of Fe(II)-based
compounds. The nature of these compounds is directly related to the corrosion mechanisms.
Moreover, the oxidation/transformation of the Fe(II)-based corrosion products may produce
other compounds (e.g., akaganeite), thus possibly leading to an erroneous interpretation of
the mechanisms.

Author Contributions: Conceptualization, all authors; methodology, all authors; validation, all authors; formal
analysis, all authors; investigation, all authors; writing—original draft preparation, P.R.; writing—review and
editing, all authors; visualization, P.R., M.J.; supervision, A.-M.G., P.R.; project administration, A.-M.G., P.R.;
funding acquisition, A.-M.G., P.R. All authors have read and agreed to the published version of the manuscript.
Funding: The various studies synthesized here were funded by: the Technological Research Institute Jules
Verne of Nantes (France) through the ADUSCOR research program; the European Regional Development Fund
(ERDF) and the CDA of La Rochelle through the DYPOMAR research project; the seaports (GPM) of La Rochelle,
Nantes-Saint-Nazaire, Marseille, and Le Havre.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Chaves, I.A.; Melchers, R.E. Long term localised corrosion of marine steel piling welds. Corros. Eng. Sci.
Technol. 2013, 48, 469–474. [CrossRef]
2. Melchers, R.E.; Jeffrey, R.J.; Usher, K.M. Localized corrosion of steel sheet piling. Corros. Sci. 2014, 79,
139–147. [CrossRef]
3. Melchers, R.E.; Chaves, I.A.; Jeffrey, R. A conceptual model for the interaction between carbon content and
manganese sulphide inclusions in the short-term seawater corrosion of low carbon steel. Metals 2016, 6, 132.
[CrossRef]
4. Liu, C.; Cheng, X.; Dai, Z.; Liu, R.; Li, Z.; Cui, L.; Chen, M.; Ke, L. Synergistic effect of Al2 O3 inclusion and
pearlite on the localized corrosion evolution process of carbon steel in marine environment. Materials 2018,
11, 2277. [CrossRef] [PubMed]
5. Liduino, V.S.; Lutterbach, M.T.S.; Sérvulo, E.F.C. Biofilm activity on corrosion of API 5L X65 steel weld bead.
Colloids Surf. B Biointerfaces 2018, 172, 43–50. [CrossRef]
6. Evans, U.R. The Corrosion and Oxidation of Metals: Scientific Principles and Practical Applications; Edward Arnold
(Publishers) Ltd.: London, UK, 1960.
7. Jeffrey, R.; Melchers, R.E. Corrosion of vertical mild steel strips in seawater. Corros. Sci. 2009, 51, 2291–2297.
[CrossRef]
8. Zou, Y.; Wang, J.; Bai, Q.; Zhang, L.L.; Peng, X.; Kong, X.F. Potential distribution characteristics of mild steel
in seawater. Corros. Sci. 2012, 57, 202–208. [CrossRef]
9. Duboscq, J.; Sabot, R.; Jeannin, M.; Refait, P.H. Localized corrosion of carbon steel in seawater: Processes
occurring in cathodic zones. Mater. Corros. 2019, 70, 973–984. [CrossRef]
10. Lee, W.; Lewandowski, Z.; Nielsen, P.H.; Hamilton, W.A. Role of sulfate-reducing bacteria in corrosion of
mild steel: A review. Biofouling 1995, 8, 165–194. [CrossRef]
11. Starovetsky, D.; Khaselev, O.; Starovetsky, J.; Armon, R.; Yahalom, J. Effect of iron exposure in SRB media on
pitting initiation. Corros. Sci. 2000, 42, 345–359.
12. Huang, Y.; Duan, J.; Ma, S. Effects of anaerobe in sea bottom sediment on the corrosion of carbon steel. Mater.
Corros. 2004, 55, 46–48. [CrossRef]
13. Jeffrey, R.; Melchers, R.E. The changing topography of corroding mild steel surfaces in seawater. Corros. Sci.
2007, 49, 2270–2288. [CrossRef]
14. Castaneda, H.; Benetton, X.D. SRB-biofilm influence in active corrosion sites formed at the steel-electrolyte
interface when exposed to artificial seawater conditions. Corros. Sci. 2008, 50, 1169–1183. [CrossRef]
Corros. Mater. Degrad. 2020, 1 216

15. Beech, I.B.; Campbell, S.A. Accelerated low water corrosion of carbon steel in the presence of a biofilm
harbouring sulphate-reducing and sulphur-oxidising bacteria recovered from a marine sediment. Electrochim.
Acta 2008, 54, 14–21. [CrossRef]
16. Dong, Z.H.; Shi, W.; Ruan, H.M.; An Zhang, G. Heterogeneous corrosion of mild steel under SRB-biofilm
characterised by electrochemical mapping technique. Corros. Sci. 2011, 53, 2978–2987. [CrossRef]
17. Melchers, R.E.; Jeffrey, R. Corrosion of long vertical steel strips in the marine tidal zone and implications for
ALWC. Corros. Sci. 2012, 65, 26–36. [CrossRef]
18. Stipaničev, M.; Turcu, F.; Esnault, L.; Schweitzer, E.W.; Kilian, R.; Basseguy, R. Corrosion behavior of carbon
steel in presence of sulfate-reducing bacteria in seawater environment. Electrochim. Acta 2013, 113, 390–406.
[CrossRef]
19. World Steel Association. “World Steel in Figures 2019”, Brussels, Belgium. 2019. Available online:
https://www.worldsteel.org/publications/infographics.html (accessed on 2 June 2020).
20. Melchers, R.E. Mathematical modelling of the diffusion controlled phase in marine immersion corrosion of
mild steel. Corros. Sci. 2003, 45, 923–940. [CrossRef]
21. Melchers, R.E.; Jeffrey, R. Early corrosion of mild steel in seawater. Corros. Sci. 2005, 47, 1678–1693. [CrossRef]
22. Melchers, R.E.; Wells, T. Models for the anaerobic phases of marine immersion corrosion. Corros. Sci. 2006,
48, 1791–1811. [CrossRef]
23. Enning, D.; Venzlaff, H.; Garrelfs, J.; Dinh, H.T.; Meyer, V.; Mayrhofer, K.J.J.; Hassel, A.W.; Stratmann, M.;
Widdel, F. Marine sulfate-reducing bacteria cause serious corrosion of iron under electroconductive biogenic
mineral crust. Environ. Microbiol. 2012, 14, 1772–1787. [CrossRef] [PubMed]
24. Venzlaff, H.; Enning, D.; Srinivasan, J.; Mayrhofer, K.J.J.; Hassel, A.W.; Widdel, F.; Stratmann, M. Accelerated
cathodic reaction in microbial corrosion of iron due to direct electron uptake by sulfate-reducing bacteria.
Corros. Sci. 2013, 66, 88–96. [CrossRef]
25. Yu, L.; Duan, J.; Du, X.; Huang, Y.; Hou, B. Accelerated anaerobic corrosion of electroactive sulfate-reducing
bacteria by electrochemical impedance spectroscopy and chronoamperometry. Electrochem. Commun. 2013,
26, 101–104. [CrossRef]
26. Pineau, S.; Sabot, R.; Quillet, L.; Jeannin, M.; Caplat, C.H.; Dupont-Morral, I.; Refait, P.H. Formation of the
Fe(II-III) hydroxysulphate green rust during marine corrosion of steel associated to molecular detection of
dissimilatory sulphite-reductase. Corros. Sci. 2008, 50, 1099–1111. [CrossRef]
27. Duan, J.; Wu, S.; Zhang, X.; Huang, G.; Du, M.; Hou, B. Corrosion of carbon steel influenced by anaerobic
biofilm in natural seawater. Electrochim. Acta 2008, 54, 22–28. [CrossRef]
28. Boudaud, N.; Coton, M.; Coton, E.; Pineau, S.; Travert, J.; Amiel, C. Biodiversity analysis by polyphasic study
of marine bacteria associated with biocorrosion phenomena. J. Appl. Microbiol. 2010, 109, 166–179. [CrossRef]
29. Usher, K.M.; Kaksonen, A.H.; MacLeod, I.D. Marine rust tubercles harbour iron corroding archaea and
sulphate reducing bacteria. Corros. Sci. 2014, 83, 189–197. [CrossRef]
30. Lanneluc, I.; Langumier, M.; Sabot, R.; Jeannin, M.; Refait, P.; Sablé, S. On the bacterial communities associated
with the corrosion product layer during the early stages of marine corrosion of carbon steel. Int. Biodeter.
Biodegrad. 2015, 99, 55–65. [CrossRef]
31. Stipanicev, M.; Turcu, F.; Esnault, L.; Rosas, O.; Basseguy, R.; Sztyler, M.; Beech, I.B. Corrosion of carbon steel
by bacteria from North Sea offshore seawater injection systems: Laboratory investigation. Bioelectrochemistry
2014, 97, 76–88. [CrossRef]
32. Rlinger, M.I.; Samokhvalov, A.A. Electron conduction in magnetite and ferrites. Phys. Stat. Sol. B 1977, 79,
9–48. [CrossRef]
33. Ihle, D.; Lorenz, B. Small-polaron band versus hopping conduction in Fe3 O4 . J. Phys. C Solid State Phys. 1985,
18, L647–L650. [CrossRef]
34. Lopez Maldonado, K.L.; de la Presa, P.; de la Rubia, M.A.; Crespo, P.; de Frutos, J.; Hernando, A.;
Matutes-Aquino, J.A.; Elizalde Galindo, J. Effects of grain boundary width and crystallite size on conductivity
and magnetic properties of magnetite nanoparticles. J. Nanopart. Res. 2014, 16, 2482. [CrossRef]
35. Li, Y.; Hou, B.; Li, H.; Zhang, J. Corrosion behavior of steel in Chengdao offshore oil exploitation area. Mater.
Corros. 2004, 55, 305–309. [CrossRef]
36. Refait, P.; Nguyen, D.D.; Jeannin, M.; Sablé, S.; Langumier, M.; Sabot, R. Electrochemical formation of green
rusts in deaerated seawater-like solutions. Electrochim. Acta 2011, 56, 6481–6488. [CrossRef]
Corros. Mater. Degrad. 2020, 1 217

37. Refait, P.; Jeannin, M.; Sabot, R.; Antony, H.; Pineau, S. Corrosion and cathodic protection of carbon steel in
the tidal zone: Products, mechanisms and kinetics. Corros. Sci. 2015, 90, 375–382. [CrossRef]
38. Morcillo, M.; Chico, B.; Alcantara, J.; Diaz, I.; Wolthuis, R.; de la Fuente, D. SEM/Micro-Raman characterization
of the morphologies of marine atmospheric corrosion products formed on mild steel. J. Electrochem. Soc.
2016, 163, C426–C439. [CrossRef]
39. Morcillo, M.; Chico, B.; de la Fuente, D.; Alcantara, J.; Odnevall Wallinder, I.; Leygraf, C. On the Mechanism
of Rust Exfoliation in Marine Environment. J. Electrochem. Soc. 2017, 164, C8–C16. [CrossRef]
40. Refait, P.; Grolleau, A.-M.; Jeannin, M.; François, E.; Sabot, R. Corrosion of carbon steel at the mud
zone/seawater interface: Mechanisms and kinetics. Corros. Sci. 2018, 130, 76–84. [CrossRef]
41. Refait, P.; Grolleau, A.-M.; Jeannin, M.; François, E.; Sabot, R. Localized corrosion of carbon steel in marine media:
Galvanic coupling and heterogeneity of the corrosion product layer. Corros. Sci. 2016, 111, 583–595. [CrossRef]
42. Refait, P.; Jeannin, M.; François, E.; Sabot, R.; Grolleau, A.-M. Galvanic corrosion in marine environments:
Effects associated with the inversion of polarity of Zn/carbon steel couples. Mater. Corros. 2019, 70, 950–961.
[CrossRef]
43. Rémazeilles, C.; Neff, D.; Kergourlay, F.; Foy, E.; Conforto, E.; Guilminot, E.; Reguer, S.; Refait, P.; Dillmann, P.
Mechanisms of long-term anaerobic corrosion of iron archaeological artefacts in seawater. Corros. Sci. 2009,
51, 2932–2941. [CrossRef]
44. Refait, P.; Jeannin, M.; Sabot, R.; Antony, H.; Pineau, S. Electrochemical formation and transformation of
corrosion products on carbon steel under cathodic protection in seawater. Corros. Sci. 2013, 71, 32–36. [CrossRef]
45. De Faria, D.L.A.; Silva, S.V.; Oliveira, M.T.D. Raman micro spectroscopy study of some iron oxides and
oxyhydroxides. J. Raman Spectrosc. 1997, 28, 873–878. [CrossRef]
46. Shebanova, O.N.; Lazor, P. Raman study of magnetite (Fe3 O4 ): Laser-induced thermal effects and oxidation.
J. Raman Spectrosc. 2003, 34, 845–852. [CrossRef]
47. Hansen, H.C.B. Composition, stabilisation, and light absorption of Fe(II)-Fe(III) hydroxycarbonate (green
rust). Clay Miner. 1989, 24, 663–669. [CrossRef]
48. Duboscq, J.; Abdelmoula, M.; Jeannin, M.; Sabot, R.; Refait, P. On the formation and transformation of
Fe(III)-containing chukanovite, FeII 2−x FeIII x (OH)2−x Ox CO3 . J. Phys. Chem. Solids 2020, 138, 109310. [CrossRef]
49. Pourbaix, M. Thermodynamics and corrosion. Corros. Sci. 1990, 30, 963–988. [CrossRef]
50. Refait, P.; Géhin, A.; Abdelmoula, M.; Génin, J.-M.R. Coprecipitation thermodynamics of iron(II-III)
hydroxysulphate green rust from Fe(II) and Fe(III) salts. Corros. Sci. 2003, 45, 656–676. [CrossRef]
51. Ruby, C.; Abdelmoula, M.; Naille, S.; Renard, A.; Khare, V.; Ona-Nguema, G.; Morin, G.; Génin, J.-M.R.
Oxidation modes and thermodynamics of FeII-III oxyhydroxycarbonate green rust: Dissolution-precipitation
versus in situ deprotonation. Geochim. Cosmochim. Acta 2010, 74, 953–966. [CrossRef]
52. Stampfl, P.P. Ein basisches eisen-II-III-karbonat in rost. Corros. Sci. 1969, 9, 185–187. [CrossRef]
53. Allmann, R. Doppelschichtstrukturen mit brucitähnlichen Schichtionen [Me(II)1−x Me(III)x (OH)2 ]x+ . Chimia
1970, 24, 99–108.
54. Refait, P.; Abdelmoula, M.; Génin, J.-M.R. Mechanisms of formation and structure of green rust one in
aqueous corrosion of iron in the presence of chloride ions. Corros. Sci. 1998, 40, 1547–1560. [CrossRef]
55. Simon, L.; François, M.; Refait, P.; Renaudin, G.; Lelaurain, M.; Génin, J.-M.R. Structure of the Fe(II-III)
layered double hydroxysulphate green rust two from Rietveld analysis. Sol. State Sci. 2003, 5, 327–334. [CrossRef]
56. Miyata, S. Anion-exchange properties of hydrotalcite-like compounds. Clays Clay Miner. 1983, 31, 305–311.
[CrossRef]
57. Mendiboure, A.; Schöllhorn, R. Formation and anion exchange reactions of layered transition metal
hydroxides [Ni1−x Mx ](OH)2 (CO3 )x/2 (H2 O)z (M=Fe,Co). Rev. Chim. Miner. 1986, 23, 819–827. [CrossRef]
58. Refait, P.; Memet, J.B.; Bon, C.; Sabot, S.; Génin, J.-M.R. Formation of the Fe(II)-Fe(III) hydroxysulphate green
rust during marine corrosion of steel. Corros. Sci. 2003, 45, 833–845. [CrossRef]
59. Bourdoiseau, J.A.; Jeannin, M.; Sabot, R.; Rémazeilles, C.; Refait, P. Characterisation of mackinawite by
Raman spectroscopy: Effects of crystallisation, drying and oxidation. Corros. Sci. 2008, 50, 3247–3255.
[CrossRef]
60. Bocher, F.; Géhin, A.; Ruby, C.; Ghanbaja, J.; Abdelmoula, M.; Génin, J.-M.R. Coprecipitation of Fe(II-III)
hydroxycarbonate green rust stabilised by phosphate adsorption. Sol. State Sci. 2004, 6, 117–124. [CrossRef]
61. Barthélémy, K.; Naille, S.; Despas, C.; Ruby, C.; Mallet, M. Carbonated ferric green rust as a new material for
efficient phosphate removal. J. Colloid Interface Sci. 2012, 384, 121–127. [CrossRef]
Corros. Mater. Degrad. 2020, 1 218

62. Usman, M.; Hanna, K.; Abdelmoula, M.; Zegeye, A.; Faure, P.; Ruby, C. Formation of green rust via
mineralogical transformation of ferric oxides (ferrihydrite, goethite and hematite). Appl. Clay Sci. 2012, 64,
38–43. [CrossRef]
63. Jeong, H.Y.; Lee, J.H.; Hayes, K.F. Characterization of nanocrystalline mackinawite: Crystal structure, particle
size, and specific surface area. Geochim. Cosmochim. Acta 2008, 72, 493–505. [CrossRef] [PubMed]
64. Ohfuji, H.; Rickard, D. High resolution transmission electron microscopic study of synthetic nanocrystalline
mackinawite. Earth Planet. Sci. Lett. 2006, 241, 227–233. [CrossRef]
65. Bourdoiseau, J.-A.; Jeannin, M.; Rémazeilles, C.; Sabot, R.; Refait, P. The transformation of mackinawite into
greigite studied by Raman spectroscopy. J. Raman Spectrosc. 2011, 42, 496–504. [CrossRef]
66. Mullet, M.; Boursiquot, S.; Abdelmoula, M.; Génin, J.-M.R.; Ehrhardt, J.J. Surface chemistry and structural
properties of mackinawite prepared by reaction of sulfide with metallic iron. Geochim. Cosmochim. Acta 2002,
66, 829–836. [CrossRef]
67. Lennie, A.R.; Redfern, S.A.T.; Champness, P.E.; Stoddart, C.P.; Schofield, P.F.; Vaughan, D.J. Transformation
of mackinawite to greigite: An in situ X-ray powder diffraction and transmission electron microscope study.
Am. Min. 1997, 82, 302–309. [CrossRef]
68. Rémazeilles, C.; Saheb, M.; Neff, D.; Guilminot, E.; Tran, K.; Bourdoiseau, J.-A.; Sabot, R.; Jeannin, M.;
Matthiesen, H.; Dillmann, P.; et al. Microbiologically influenced corrosion of archaeological artefacts;
characterisation of iron(II) sulphides by Raman spectroscopy. J. Raman Spectrosc. 2010, 41, 1135–1143.
[CrossRef]
69. Malard, E.; Kervadec, D.; Gil, O.; Lefevre, Y.; Malard, S. Interactions between steels and sulphide-producing
bacteria-Corrosion of carbon steels and low-alloy steels in natural seawater. Electrochim. Acta 2008, 54, 8–13.
[CrossRef]
70. Refait, P.; Sabot, R.; Jeannin, M. Role of Al(III) and Cr(III) on the formation and oxidation of the Fe(II-III)
hydroxysulfate Green Rust. Colloids Surf. A Phys. Eng. Asp. 2017, 531, 203–212. [CrossRef]
71. Hartt, W.H.; Culberson, C.H.; Smith, S.W. Calcareous deposits on metal surfaces in seawater- A critical
review. Corrosion 1984, 40, 609–618. [CrossRef]
72. Lee, R.U.; Ambrose, J.R. Influence of cathodic protection parameters on calcareous deposit formation.
Corrosion 1986, 44, 887–891. [CrossRef]
73. Yan, J.F.; White, R.E.; Griffin, R.B. Parametric studies of the formation of calcareous deposits on cathodically
protected steel in seawater. J. Electrochem. Soc. 1993, 141, 1275–1280. [CrossRef]
74. Barchiche, C.; Deslouis, C.; Festy, D.; Gil, O.; Refait, P.; Touzain, S.; Tribollet, B. Characterization of calcareous
deposits in artificial sea water by impedance techniques. 3- Deposit of CaCO3 in the presence of Mg(II).
Electrochim. Acta 2003, 48, 1645–1654. [CrossRef]
75. Cook, D.C.; van Orden, A.C.; Carpio, J.J.; Oh, S.J. Atmospheric corrosion in the Gulf of Mexico. Hyperf.
Interact. 1998, 113, 319–329. [CrossRef]
76. Li, S.; Hihara, L.H. A micro-Raman spectroscopic study of marine atmospheric corrosion of carbon steel: The
effect of akaganeite. J. Electrochem. Soc. 2015, 162, C495–C502. [CrossRef]
77. Stahl, K.; Nielsen, K.; Jiang, J.; Lebech, B.; Hanson, J.C.; Norby, P.; van Lanschot, J. On the akaganeite crystal
structure, phase transformations and possible role in post-excavational corrosion of iron artifacts. Corros. Sci.
2003, 45, 2563–2575. [CrossRef]
78. Rémazeilles, C.; Refait, P. On the formation of β-FeOOH (Akaganéite) in chloride-containing environments.
Corros. Sci. 2007, 49, 844–857. [CrossRef]
79. Refait, P.; Génin, J.-M.R. The mechanisms of oxidation of ferrous hydroxychloride β-Fe2 (OH)3 Cl in
chloride-containing aqueous solution: The formation of β-FeOOH akaganeite; an X-ray diffraction, Mössbauer
spectroscopy and electrochemical study. Corros. Sci. 1997, 39, 539–553. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like